• Aucun résultat trouvé

Maximal Regularity for Non-Autonomous Second Order Cauchy Problems

N/A
N/A
Protected

Academic year: 2021

Partager "Maximal Regularity for Non-Autonomous Second Order Cauchy Problems"

Copied!
23
0
0

Texte intégral

(1)

HAL Id: hal-00880886

https://hal.archives-ouvertes.fr/hal-00880886

Submitted on 7 Nov 2013

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

Maximal Regularity for Non-Autonomous Second Order

Cauchy Problems

Dominik Dier, El Maati Ouhabaz

To cite this version:

Dominik Dier, El Maati Ouhabaz. Maximal Regularity for Non-Autonomous Second Order Cauchy

Problems. Integr. Eqs and Op. Theory, 2014, 78 (3), pp.427-450. �10.1007/s00020-013-2109-6�.

�hal-00880886�

(2)

Maximal Regularity for Non-Autonomous Second

Order Cauchy Problems

Dominik Dier, El Maati Ouhabaz

November 8, 2013

Abstract

We consider non-autonomous wave equations

u¨(t) + B(t) ˙u(t) + A(t)u(t) = f (t) t-a.e.

u(0) = u0, ˙u(0) = u1.

where the operators A(t) and B(t) are associated with time-dependent sesquilinear forms a(t, ., .) and b defined on a Hilbert space H with the same domain V . The initial values satisfy u0∈ V and u1∈ H. We prove

well-posedness and maximal regularity for the solution both in the spaces

Vand H. We apply the results to non-autonomous Robin-boundary

conditions and also use maximal regularity to solve a quasilinear problem.

Key words: Sesquilinear forms, non-autonomous evolution equations, maximal regularity, non-linear heat equations, wave equation.

MSC:35K90, 35K45, 35K92, 47F05.

1

Introduction

The present paper is a continuation of [ADLO13] which is devoted to maximal regularity for first order non-autonomous evolution equations governed by forms. Here we address the problem of maximal regularity for non-autonomous second order problems.

We consider Hilbert spaces H and V such that V is continuously embedded into

H and two families of sesquilinear forms

a: [0, T ] × V × V → C, b: [0, T ] × V × V → C

such that a(., u, v) : [0, T ] → C, b(., u, v): [0, T ] → C are measurable for all

u, v∈ V ,

|a(t, u, v)| ≤ MkukVkvkV (t ∈ [0, T ]),

and

Re a(t, u, u) + wkuk2H ≥ αkuk2V (u ∈ V, t ∈ [0, T ])Corresponding author.

(3)

where M ≥ 0, w ∈ R, and α > 0 are constants. We assume also that b satisfies the same properties. For fixed t ∈ [0, T ], we denote by A(t), B(t) ∈ L(V, V

) the operators associated with the forms a(t, ., .) and b(t, ., .), respectively. Given a function f defined on [0, T ] with values either in H or in Vand consider the

second order evolution equation (

¨

u(t) + B(t) ˙u(t) + A(t)u(t) = f(t) t-a.e.

u(0) = u0, ˙u(0) = u1.

(1.1)

with initial values u0∈ V and u1∈ H. This is a damped non-autonomous wave

equation. The equation without the factor ˙u, i.e., (

¨

u(t) + A(t)u(t) = f(t) t-a.e.

u(0) = u0, ˙u(0) = u1.

(1.2)

is a non-autonomous wave equation.

Our aim is to prove well-posedness and maximal regularity for (1.1) and (1.2). We shall prove three main results. The first one concerns maximal reg-ularity in Vfor the damped wave equation (1.1). We prove that for u

0 ∈ V, u1∈ H and f ∈ L2(0, T, V) there exists a unique solution u ∈ H1(0, T, V ) ∩ H2(0, T, V). This result was first proved by Lions [Lio61, p. 151] by assuming

regularity of t 7→ a(t, u, v) and t 7→ b(t, u, v) for every fixed u, v ∈ V . This regularity assumption was removed in Dautray-Lions [DL88, p. 667], but taking

f ∈ L2(0, T, H) and considering mainly symmetric forms. The general case was

given recently by Batty, Chill and Srivastava [BCS08] by reducing the problem to a first order non-autonomous equation. The result in [BCS08] is stated in the case u0= u1= 0, only. Our proof is different from [BCS08] and is inspired by

that of Lions [Lio61]. Next we consider maximal regularity in H. This is more delicate and needs extra properties on the forms a and b. We prove that if the forms are symmetric and t 7→ a(t, u, v) and t 7→ b(t, u, v) are piecewise Lipschitz on [0, T ] then for u0 ∈ V , u1 ∈ H and f ∈ L2(0, T, H) there exists a unique

solution u ∈ H1(0, T, V )∩H2(0, T, H) to the equation (1.1). We also allow some

non-symmetric perturbations of a and b. The third result (Theorem 5.1) con-cerns the wave equation (1.2). We prove that if a is symmetric and t 7→ a(t, u, v) is Lipschitz on [0, T ], then for every u0∈ V , u1 ∈ H and f ∈ L2(0, T, H) there

exists a unique solution u ∈ L2(0, T, V )∩H1(0, T, H)∩H2(0, T, V) to the

equa-tion (1.2). This result is not new and was already proved by Lions [Lio61, p. 150] for the case u0= 0 and later in [DL88, p. 666] for u0∈ V and u1∈ H.

The-orem 5.1 is stated in order to have a complete picture of maximal regularity for wave equations with or without damping. The proof in [DL88] uses a Galerkin method and sectorial approximation. The proofs of the three main theorems use a representation result of Lions (see Theorem 2.4 below) for a given sesquilinear form E acting on a product of a Hilbert and pre-Hilbert spaces H × V. In each case we have to define the appropriate spaces H, V and the form E to which we apply Theorem 2.4. This idea was already used in [Lio61]. Our choice of the spaces H, V and the form E allow us to sharpen and extend some results from [Lio61] and assume less regularity on t 7→ a(t, u, v) and t 7→ b(t, u, v).

We illustrate our abstract results by two examples. The first one is a linear damped wave equation with time dependent Robin boundary conditions. The second is a quasi-linear second order non-autonomous problem. The latter is

(4)

treated by a fixed point argument but the implementation of this classical idea uses heavily a priori estimates that follow from our maximal regularity results for linear equations.

Acknowledgment

Some ideas in this work germinated during a visit of the second named author at the University of Ulm in the framework of the Graduate School: Mathematical Analysis of Evolution, Information and Complexity financed by the Land Baden-Württemberg and during the visit of the first named author at the University of Bordeaux. Both authors thank Wolfgang Arendt for fruitful discussions on the non-autonomous maximal regularity.

D. Dier is a member of the DFG Graduate School 1100: Modeling, Analysis and Simulation in Economics.

The research of E. M. Ouhabaz is partly supported by the ANR project “Har-monic Analysis at its Boundaries”, ANR-12-BS01-0013-02.

2

Preliminaries

Throughout this paper, V and H are separable Hilbert spaces over the field K= C or R. The scalar products of H and V and the corresponding norms will be denoted by (. | .)H, (. | .)V, k.kHand k.kV, respectively. We denote by V′ the

antidual of V when K = C and the dual when K = R. The duality between V

and V is denoted by h., .i. Then hu, vi = (u | v)H for u ∈ H and v ∈ V .

We assume that

V ֒

d H;

i.e., V is a dense subspace of H such that for some constant cH>0,

kukH≤ cHkukV (u ∈ V ). (2.1)

By duality and density of V in H one has

H ֒

d V

.

The space H is then identified with a dense subspace of V

(associating to u ∈ H the antilinear map v 7→ (u | v)H = hu, vi for v ∈ V ).

Let

a: [0, T ] × V × V → K be a family of sesquilinear and V -bounded forms; i.e.

|a(t, u, v)| ≤ MkukVkvkV (u, v ∈ V, t ∈ [0, T ]) (2.2)

for some constant M , such that a(., u, v) is measurable for all u, v ∈ V . We shall call a satisfying the above properties a V -bounded non-autonomous sesquilinear

form. Moreover we say that a is quasi-coercive if there exist constants α > 0, ω∈ R such that

Re a(t, u, u) + ωkuk2H≥ αkuk

2

V (u ∈ V, t ∈ [0, T ]). (2.3)

(5)

For t ∈ [0, T ], a V -bounded and quasi-coercive sesquilinear form a(t, ., .) is closed. The operator A(t) ∈ L(V, V

) associated with a(t, ., .) is defined by hA(t)u, vi = a(t, u, v) for u, v ∈ V. (2.4) We may also associate with a(t, ., .) an operator on H by taking the part A(t) of A(t) on H; i.e.,

D(A(t)) := {u ∈ V : A(t)u ∈ H}

A(t)u := A(t)u. Note that if a(t, ., .) is symmetric, i.e.,

a(t, u, v) = a(t, v, u) for all u, v ∈ V , then the operator A(t) is self-adjoint.

For a Hilbert space E we denote by L2(0, T, E) the L2-space on (0, T ) of

functions with values in E and by Hk(0, T, E) we denote the usual Sobolev

space of order k of functions on (0, T ) with values in E. For u ∈ H1(0, T ; E) we

denote the first derivative by ˙u and for u ∈ H2(0, T ; E) the second derivative

by ¨u.

We start with the following differentiation result.

Lemma 2.1. Let

a: [0, T ] × V × V → K

be a V -bounded, quasi-coercive non-autonomous form. Suppose that it is Lips-chitz with LipsLips-chitz constant ˙M, that is

|a(t, φ, ψ) − a(s, φ, ψ)| ≤ ˙M|t − s|kφkVkψkV, t, s∈ [0, T ] and φ, ψ ∈ V.

Let u, v ∈ H1(0, T ; V ). Then a(., u, v) ∈ W1,1(0, T ) and there exists a non-autonomous form ˙a which is V -bounded with constant ˙M such that

a(., u, v)˙ = a(., u, ˙v) + a(., ˙u, v) + ˙a(., u, v)

If additionally a is symmetric then

a(., u, u)˙ = 2 Re a(., u, ˙u) + ˙a(., u, u).

Note that for u, v ∈ V we have dtda(t, u, v) = ˙a(t, u, v) for a.e. t ∈ [0, T ].

This lemma is a consequence of the next two results.

Lemma 2.2. Let u ∈ H1(0, T ; V ) and v ∈ H1(0, T ; V

). Then hv(.), u(.)i ∈

W1,1(0, T ) and

hv(.), u(.)i˙ = h ˙v(.), u(.)i + hv(.), ˙u(.)i.

Proof. By Fubini’s Theorem we have

Z t 0 h ˙v(s), u(s)i ds = Z t 0 D ˙v(s), u(0) + Z s 0 ˙u(r) drEds = hv(t), u(0)i − hv(0), u(0)i + Z t 0 Z s 0 h ˙v(s), ˙u(r)i dr ds

(6)

= hv(t), u(0)i − hv(0), u(0)i + Z t 0 Z t r h ˙v(s), ˙u(r)i ds dr = hv(t), u(0)i − hv(0), u(0)i + Z t 0 hv(t), ˙u(r)i − hv(r), ˙u(r)i dr = hv(t), u(t)i − hv(0), u(0)i − Z t 0 hv(r), ˙u(r)i dr. Thus hv(t), u(t)i = hv(0), u(0)i + Z t 0h ˙v(s), u(s)i ds + Z t 0 hv(s), ˙u(s)i ds

which proves the claim.

Proposition 2.3. Let S : [0, T ] → L(V, V) be Lipschitz continuous. Then the following assertions hold.

a) There exists a bounded, strongly measurable function ˙S: [0, T ] → L(V, V

)

such that

d

dtS(t)u = ˙S(t)u (u ∈ V )

for a.e. t ∈ [0, T ] and

k ˙S(t)kL(V,V′)≤ L (t ∈ [0, T ]) where L is the Lipschitz constant of S.

b) If u ∈ H1(0, T ; V ), then Su := S(.)u(.) ∈ H1(0, T ; V) and

(Su)˙ = ˙S(.)u(.) + S(.) ˙u(.). (2.5) Proposition 2.3 is proved in [ADLO13]. Lemma 2.1 follows from (2.4), Lemma 2.2 and Proposition 2.3.

We shall need the following representation result due to Lions. See [Lio59, p. 156], [Lio61, p. 61] or [ADLO13].

Theorem 2.4 (Lions’ Representation Theorem). Let H be a Hilbert space, V

a pre-Hilbert space such that V ֒→ H. Let E : H × V → K be sesquilinear such that

1) for all w ∈ V, E(., w) is a continuous linear functional on H; 2) |E(w, w)| ≥ αkwk2V for all w ∈ V

for some α > 0. Let L ∈ V

. Then there exists u ∈ H such that Lw= E(u, w)

(7)

3

Maximal Regularity for the Damped Wave

Equation in V

Let H, V be Hilbert spaces such that V ֒→ H. We define the following maximald

regularity space

MR(V, V, V) := L2(0, T, V ) ∩ H1(0, T ; V ) ∩ H2(0, T ; V′) = H1(0, T ; V ) ∩ H2(0, T ; V).

Let a : [0, T ] × V × V → C and b: [0, T ] × V × V → C be non-autonomous

V-bounded and quasi-coercive sesquilinear forms. We denote by A(t) and B(t)

their associated operators in the sense of (2.4). The following is our first result.

Theorem 3.1. For every u0 ∈ V , u1∈ H and f ∈ L2(0, T ; V), there exists a unique solution u ∈ MR(V, V, V) of the non-autonomous second order Cauchy problem

( ¨

u(t) + B(t) ˙u(t) + A(t)u(t) = f(t) t-a.e.

u(0) = u0, ˙u(0) = u1.

(3.1)

Moreover there exists a constant C > 0 such that

kukMR(V,V,V′)≤ C

h

ku0kV + ku1kH+ kfkL2(0,T ;V)

i

. (3.2) As mentioned in the introduction, this theorem was first proved by Lions [Lio61, p. 151] under an additional regularity assumption on t 7→ a(t, u, v) and

t7→ b(t, u, v). This regularity assumption was removed in Dautray-Lions [DL88, p. 667], but taking f ∈ L2(0, T, H) and considering mainly symmetric forms

(they allow some non-symmetric perturbations). Their proof is based on a Galerkin method. Another proof of Theorem 3.1 was given recently by Batty, Chill and Srivastava [BCS08] but they consider only the case u0= u1= 0. Our

proof is based on Theorem 2.4 and is in the spirit of Lions [Lio61]. It is different from the proofs in [DL88] and [BCS08].

A classical result of Lions says that

MR(V, V

) := L2(0, T, V ) ∩ H1(0, T ; V

) ֒→ C([0, T ]; H), (3.3) and also that for u ∈ MR(V, V

) the function ku(.)k2H is in W1,1(0, T ) with

(kuk2H)˙ = 2 Reh ˙u, ui, (3.4)

see [Sho97, p. 106] and [DL88, p.570]. This implies that MR(V, V, V

) ֒→

C([0, T ]; V ) ∩ C1([0, T ]; H). Thus for u ∈ MR(V, V, V), both u(0) and ˙u(0)

make sense.

We start with the following basic lemma.

Lemma 3.2. For v ∈ H1(0, T ; V ) we have Z T 0 kv(t)k 2 V dt 1/2 ≤ T Z T 0 k ˙v(s)k 2 V ds 1/2 +√Tkv(0)kV.

(8)

Proof. Note that v(t) = v(0) +Rt 0 ˙v(s) ds, thus Z T 0 kv(t)k 2 V dt = Z T 0  v(0) + Z t 0 ˙v(s) ds v(t)  V dt = Z T 0 Z T s ( ˙v(s) | v(t)) V dt ds + Z T 0 (v(0) | v(t)) V dt ≤ Z T 0 Z T 0 k ˙v(s)kVkv(t)kV dt ds + Z T 0 kv(0)kVkv(t)kV dt ≤ Z T 0 kv(t)kV dt Z T 0 k ˙v(s)kV ds + kv(0)kV  ≤T Z T 0 kv(t)k 2 V dt 1/2 T Z T 0 k ˙v(s)k 2 V ds 1/2 + kv(0)kV  .

Proof of Theorem 3.1. It suffices to show that there exists a unique solution in

the case where T < T0 and T0 > 0 is a constant that depends only on the

constants M, ω and α of (2.2) and (2.3). Indeed we can extend this solution to [0, T ] for any fixed T as follows. We write the interval [0, T ] as a finite union of sub-intervals [τi, τi+1], each has length less than T0. On each interval

[τi, τi+1] we have a unique solution ui with ui(τi) ∈ V , ˙ui(τi) ∈ H and ui

MR(V, V, V

) ֒→ C1([τ

i, τi+1]; H) ∩ C([τi, τi+1]; V ). On [τi+1, τi+2] we solve the

equation with ui+1

i+1) = ui(τi+1) and ˙ui+1(τi+1) = ˙ui(τi+1). We define u on

[0, T ] by u = ui on [τ

i, τi+1] and check easily that u ∈ MR(V, V, V) (on [0, T ])

is the unique solution to (3.1).

We prove existence of a solution in the case where

T < T0= min  α2 M2, α2M  . (3.5)

Note that we may assume throughout this proof that the forms a and b are both coercive. Indeed, set v(t) = ewtu(t) then we have

¨

v(t) + B(t) ˙v(t) + A(t)v(t)

= ewt ¨u(t) + (B(t) + 2w) ˙u(t) + (A(t) + wB(t) + w2)u(t) . (3.6)

Since a and b are quasi-coercive, we may choose w large enough such that b+2w and a + wb + w2 are coercive. Note also that v ∈ MR(V, V, V) if and only if u∈ MR(V, V, V).

We define the Hilbert space H := H1(0, T ; V ) endowed with its usual norm

kukH:= kukH1(0,T ;V ) and the pre-Hilbert space

V := {v ∈ H2(0, T ; V ) : ˙v(T ) = 0}

with norm k.kV:= k.kH. Further we define the sesquilinear form E : H × V → C

by E(u, v) := − Z T 0 ( ˙u | ¨v) H dt + Z T 0 b(t, ˙u, ˙v) dt + Z T 0 a(t, u, ˙v) dt + a(0, u(0), v(0))

(9)

and for u0∈ V , u1∈ H and f ∈ L2(0, T ; V) we define F : V → C by F(v) := Z T 0 hf, ˙vi dt + a(0, u 0, v(0)) + (u1| ˙v(0))H. We claim that 1) E(., v) ∈ Hand F ∈ V;

2) E is coercive; i.e., there exists a C > 0 such that |E(v, v)| ≥ Ckvk2H for

all v ∈ V.

Suppose for a moment that 1) and 2) are satisfied. Then we can apply Lions’s representation theorem (see Theorem 2.4) and obtain u ∈ H such that

E(u, v) = F (v) ∀ v ∈ V. (3.7) We show that u is a solution of (3.1).

Let ψ(t) ∈ D(0, T ) and w ∈ V and choose v(t) :=Rt

0ψ(s) ds w. It follows from (3.7) that − Z T 0 h ˙u(t), wi ˙ ψ(t)dt = Z T

0 hf(t) − B(t) ˙u − A(t)u(t), wiψ(t) dt.

This means that ˙u ∈ H1(0, T ; V

), hence u ∈ MR(V, V, V) and

¨

u(t) + B(t) ˙u(t) + A(t)u(t) = f(t) t-a.e. (3.8) in V

. For general v ∈ V, we use again (3.7) and integration by parts to obtain

( ˙u(0) | ˙v(0))H+ Z T 0 h¨u, ˙vi dt + Z T 0 b(t, ˙u, ˙v) dt + Z T 0 a(t, u, ˙v) dt + a(0, u(0), v(0)) = Z T 0 hf, ˙vi dt + a(0, u 0, v(0)) + (u1| ˙v(0))H.

This equality together with (3.8) imply that

( ˙u(0) | ˙v(0))H+ a(0, u(0), v(0)) = a(0, u0, v(0)) + (u1| ˙v(0))H.

Since v ∈ V is arbitrary we obtain that u(0) = u0and ˙u(0) = u1. Therefore, u is

a solution of (3.1) on [0, T ] for T ≤ T0 and T0 is such that the above properties

1) and 2) are satisfied.

Now we return to 1) and 2). Property 1) is obvious. We show the coercivity property 2). Let v ∈ V. The equality d

dtk ˙v(t)k 2 H = 2 Re(¨v(t) | ˙v(t))H implies Z T 0 Re(¨v| ˙v)H dt = −1 2k ˙v(0)k 2 H. It follows that |E(v, v)| ≥ Re E(v, v) = 1 2k ˙v(0)k 2 H+ Z T 0 Re b(t, ˙v, ˙v) dt

(10)

+ Z T

0

Re a(t, v, ˙v) dt + Re a(0, v(0), v(0)). We use coercivity of b, a and V -boundedness of a to obtain

|E(v, v)| ≥ 12k ˙v(0)k2H+ α Z T 0 k ˙vk 2 V dt − M Z T 0 kvkVk ˙vkV dt + αkv(0)k 2 V.

Therefore, by Young’s inequality, we have

|E(v, v)| ≥ 12k ˙v(0)k2H+ α 2 Z T 0 k ˙vk 2 V dt − M2 Z T 0 kvk 2 V dt + αkv(0)k 2 V.

Next we apply Lemma 3.2 to obtain

|E(v, v)| ≥ α2M 2T2 α  Z T 0 k ˙vk 2 V dt +  αM 2T α  kv(0)k2V.

Now we use (3.5) and the fact that by Lemma 3.2,RT

0 kvk 2 V dt is dominated (up to a constant) byRT 0 k ˙vk 2 V dt + kv(0)k 2 V. We obtain 2).

Next we prove uniqueness. Suppose that u and v are two solutions of (3.1) which are in MR(V, V, V

). Set w = u − v. Clearly w ∈ MR(V, V, V) and

satisfies (in V)

¨

w(t) + B(t) ˙w(t) + A(t)w(t) = 0, w(0) = 0, ˙w(0) = 0. We show that w = 0. For fixed t ∈ (0, T ] we have

Z t 0 Reh ¨ w,w˙i ds + Z t 0 Re b(s, ˙w,w˙) ds + Z t 0 Re a(s, w, ˙w) ds = 0. Using (3.4) we have Z t 0 Reh ¨ w,w˙i ds = 1 2 Z t 0 k ˙ wk2H˙ ds = 1 2k ˙w(t)k 2 H− 1 2k ˙w(0)k 2 H= 1 2k ˙w(t)k 2 H, and hence 0 = 1 2k ˙w(t)k 2 H+ Z t 0 Re b(s, ˙w,w˙) ds + Z t 0 Re a(s, w, ˙w) ds ≥ 12k ˙w(t)k2H+ α Z t 0 k ˙ wk2V ds − M Z t 0 kwkVk ˙ wkV ds.

Here we used coercivity of b and V -boundedness of a. Therefore, by Lemma 3.2, we have 0 ≥ 12k ˙w(t)k2H+ α Z t 0 k ˙ wk2V ds − M Z t 0 kwk 2 V ds 1/2Z t 0 k ˙ wk2V ds1/2 ≥ 12k ˙w(t)k2H+ (α − MT ) Z t 0 k ˙ wk2V ds.

(11)

By (3.5) we obtain that w = 0. This shows uniqueness.

Finally, in order to prove the apriori estimate (3.2), we consider the operator

S: V × H × L2(0, T, V) 7→ MR(V, V, V), (u0, u1, f) 7→ u.

This is a linear operator which is well defined thanks to the uniqueness of the solution u of (3.1). It is easy to see that S is a closed operator. Therefore it is continuous by the closed graph theorem. This gives (3.2).

The previous proof does not give any information on the constant C in (3.2). For small time T one can prove that C depends only on the constants of the forms. This observation will be needed in our application to a quasi-linear problem.

Proposition 3.3. If T > 0 is small enough, then the constant C in (3.2) depends only on the constants w, α, M and T .

Proof. Let u ∈ MR(V, V, V

) be the solution of (3.1). For fixed t ∈ (0, T ] we have Z t 0 Rehf, ˙ui ds = Z t 0 Reh¨u, ˙ui ds + Z t 0 Re b(s, ˙u, ˙u) ds + Z t 0 Re a(s, u, ˙u) ds. Since by (3.4) Z t 0 Reh¨u, ˙ui ds = 1 2 Z t 0 k ˙uk 2 H˙ ds = 1 2k ˙u(t)k 2 H− 1 2k ˙u(0)k 2 H,

it follows by Young’s inequality that 1 α Z t 0 kfk 2 Vds + α 4 Z t 0 k ˙uk 2 V ds ≥ Z t 0 kfkVk ˙ukV ds ≥ Z t 0 Rehf, ˙ui ds =1 2k ˙u(t)k 2 H− 1 2k ˙u(0)k 2 H+ Z t 0 Re b(s, ˙u, ˙u) ds + Z t 0 Re a(s, u, ˙u) ds ≥ −12k ˙u(0)k2H+ α Z t 0 k ˙uk 2 V ds − M Z t 0 kukVk ˙ukV ds ≥ −1 2k ˙u(0)k 2 H+ 4 Z t 0 k ˙uk 2 V ds − M2 α Z t 0 kuk 2 V ds.

Here we used coercivity of b and V -boundedness of a. Therefore, by Lemma 3.2, we have 1 α Z t 0 kfk 2 Vds + 1 2k ˙u(0)k 2 Hα 2 Z t 0 k ˙uk 2 V ds − M2 α Z t 0kuk 2 V ds α 2 − t2(2M2+ α) α  Z t 0 k ˙uk 2 V ds − t 2M 2+ α α  ku(0)k2V + 1 2 Z t 0 kuk 2 V ds. (3.9)

where we choose t such that α

2 > t2 (2M2 +α) α . Finally, since ¨

(12)

we obtain that k¨u(s)k2V≤ 3kf(s)k 2 V+ 3M k ˙u(s)k 2 V + 3M ku(s)k 2 V s-a.e.

This together with (3.9) ends the proof of the proposition when T is such that

α

2 >

T2(2M2+ α)

α .

4

Maximal Regularity for the Damped Wave

Equation in H

Let V, H be separable Hilbert spaces such that V ֒→

d H and let

a, b: [0, T ] × V × V → K

be closed non-autonomous sesquilinear forms on which we impose the following conditions. Each can be written as the sum of two non-autonomous forms a(t, u, v) = a1(t, u, v) + a2(t, u, v), b(t, u, v) = b1(t, u, v) + b2(t, u, v) u, v ∈ V where

a1, b1: [0, T ] × V × V → K satisfy the following assumptions

a) |a1(t, u, v)| ≤ MkukVkvkV for all u, v ∈ V , t ∈ [0, T ];

b) a1(t, u, u) ≥ αkuk2V for all u ∈ V , t ∈ [0, T ] with α > 0;

c) a1(t, u, v) = a1(t, v, u) for all u, v ∈ V , t ∈ [0, T ];

d) a1 is piecewise Lipschitz-continuous; i.e., there exist 0 = τ0 < τ1 <· · · < τn= T such that

|a1(t, u, v) − a1(s, u, v)| ≤ ˙M|t − s|kukVkvkV

for all u, v ∈ V, s, t ∈ [τi−1, τi], i ∈ {1, . . . , n},

and similarly for b1. Of course we may choose the same constants M, ˙M and α

for both forms a1and b1. We may also choose that same sub-intervals 0 = τ0< τ1<· · · < τn= T for both forms.

The non-autonomous forms

a2, b2: [0, T ] × V × V → K are measurable and satisfy

(13)

and similarly for b2.

Note that by Lemma 2.1, if c is a Lipschitz form on [0, T ], we may define its derivative ˙c(t, ., .) and we have

|˙c(t, u, v)| ≤ ˙MkukVkvkV, u, v∈ V (4.1)

for some constant ˙M. We shall use this estimate for c = a1 and for c = b1 on

sub-intervals of [0, T ] where these forms are supposed to be Lipschitz.

Let us denote by A(t) and B(t) the operators given by hA(t)u, vi = a(t, u, v) and hB(t)u, vi = b(t, u, v) for all u, v ∈ V .

As in the previous section we consider the damped wave equation. Here we study the maximal regularity property in H rather than in V

. We introduce the maximal regularity space

MR(V, V, H) := H1(0, T ; V ) ∩ H2(0, T ; H).

We have

Theorem 4.1. Let a = a1+ a2and b = b1+ b2 be non-autonomous V -bounded and quasi-coercive forms satisfying the above properties a) − e). Then for every u0, u1∈ V and f ∈ L2(0, T ; H), there exists a unique solution u ∈ MR(V, V, H) of the non-autonomous second order Cauchy problem

( ¨

u(t) + B(t) ˙u(t) + A(t)u(t) = f(t) t-a.e.

u(0) = u0,˙u(0) = u1

(4.2)

Moreover ˙u(t) ∈ V for all t ∈ [0, T ].

For a related result see Lions [Lio61, p. 155]. However the result proved there is restricted to u1 = 0 and assumes f, f∈ L2(0, T, H). Our proof resembles

that of Theorem 3.1 and uses similar ideas as in Lions [Lio61]. We use the following lemma for the proof of Theorem 4.1.

Lemma 4.2. Suppose that the forms a1 and b1 are Lipschitz continuous on

[0, T ]. Let v ∈ H2(0, T ; V ) and ǫ > 0. Then

(i) Z T 0 eλtRe b1(t, ˙v, ¨v) dt = λ2 Z T 0 eλtb1(t, ˙v, ˙v) dt −1 2 Z T 0 eλt˙b1(t, ˙v, ˙v) dt +12eλTb1(T, ˙v(T ), ˙v(T )) − 1 2b1(0, ˙v(0), ˙v(0)). (ii) Z T 0 eλtRe a 1(t, v, ¨v) dt = λ2eλTa1(T, v(T ), v(T )) −λ2a1(0, v(0), v(0)) + eλTRe a1(T, v(T ), ˙v(T )) − Re a1(0, v(0), ˙v(0)) +λ22 Z T 0 eλta1(t, v, v) dt −λ 2 Z T 0 eλt˙a1(t, v, v) dt − Z T 0 eλtRe ˙a1(t, v, ˙v) dt − Z T 0 eλta1(t, ˙v, ˙v) dt. (iii) Z T 0 eλtRe b 1(t, ˙v, ¨v) dt + Z T 0 eλtRe a 1(t, v, ¨v) dt ≥ 1 2(αλ − 2 ˙M− 2M) Z T 0 eλtk ˙vk2V dt

(14)

+ (λ2(αλ − ˙M) −M˙22) Z T 0 eλtkvk2V dt +1 2eλTh (α − ǫ)k ˙v(T )k2V + (λα − ˙ M2 ǫ )kv(T )k 2 V i −12b1(0, ˙v(0), ˙v(0)) − λ 2a1(0, v(0), v(0)) − Re a1(0, v(0), ˙v(0)). Proof. The proof of (i) and (ii) is based on Lemma 2.1 and the product rule.

Part (i) is a direct consequence of the formulae

eλtb1(t, ˙v, ˙v)˙ = −λeλtb1(t, ˙v, ˙v) + eλt˙b1(t, ˙v, ˙v) + 2eλtRe b1(t, ˙v, ¨v). For (ii) we first calculate the following derivatives

Re eλta1(t, v, v)˙ = −λeλta1(t, v, v) + eλtRe ˙a1(t, v, v) + 2eλtRe a 1(t, v, ˙v) Re eλta 1(t, v, ˙v)˙ = −λeλtRe a1(t, v, ˙v) + eλtRe ˙a1(t, v, ˙v) + eλta 1(t, ˙v, ˙v) + eλtRe a1(t, v, ¨v)

then we multiply the first equation by λ

2 and add the second equation. Now (ii)

follows by integration over t from 0 to T .

For (iii) we add (i) and (ii) and use coercivity of a1, b1 and V -boundedness

of a1, ˙a1, b1, ˙b1. Thus Z T 0 eλtRe b1(t, ˙v, ¨v) dt + Z T 0 eλtRe a1(t, v, ¨v) dt ≥ 12(αλ − ˙M− 2M) Z T 0 eλt k ˙vk2V dt +λ 2(αλ − ˙M) Z T 0 eλtkvk2V dt − ˙M Z T 0 eλtkvkVk ˙vkV dt +12eλThα k ˙v(T )k2V + λαkv(T )k 2 V − 2Mkv(T )kVk ˙v(T )kV i −12b1(0, ˙v(0), ˙v(0)) − λ 2a1(0, v(0), v(0)) − Re a1(0, v(0), ˙v(0)).

We apply Young’s inequality and see that the last term is bounded from below by 1 2(αλ − 2 ˙M− 2M) Z T 0 eλtk ˙vk2V dt + (λ2(αλ − ˙M) − ˙ M2 2 ) Z T 0 eλtkvk2V dt +1 2eλTh (α − ǫ)k ˙v(T )k2V + (λα − ˙ M2 ǫ )kv(T )k 2 V i −12b1(0, ˙v(0), ˙v(0)) −λ2a1(0, v(0), v(0)) − Re a1(0, v(0), ˙v(0)) for ǫ > 0.

Proof of Theorem 4.1. Uniqueness follows from Theorem 3.1 and we only need

to prove existence of a solution. As in the proof of Theorem 3.1 we may assume that the forms a and b are both coercive (see (3.6)).

1- Lipschitz-continuous forms. Suppose first that the forms a1 and b1 are

(15)

We define the Hilbert space

H := {u ∈ H2(0, T ; H) ∩ H1(0, T ; V ) : u(0), ˙u(0), ˙u(T ) ∈ V } with norm kukH given by

kuk2H:= k¨uk 2 L2 (0,T ;H)+ kuk 2 H1 (0,T ;V )+ ku(0)k 2 V + k ˙u(0)k 2 V + k ˙u(T )k 2 V

and the pre-Hilbert space V := H2(0, T ; V ) with norm k.k

V := k.kH. Next we

define the sesquilinear form E : H × V → C by

E(u, v) := Z T 0 eλtu | ¨v)H dt + Z T 0 eλtb(t, ˙u, ¨v) dt +Z T 0 eλta(t, u, ¨v) dt + η( ˙u(0) | ˙v(0))V + η(u(0) | v(0))V,

where λ and η are positive parameters. Later on, we will choose them to be large enough. For u0, u1∈ V and f ∈ L2(0, T ; H), we define F : V → C by

F(v) := Z T

0 eλt

(f | ¨v)H dt + η(u1| ˙v(0))V + η(u0| v(0))V

We proceed as in the proof of Theorem 3.1. Suppose for a moment that 1) E(., v) ∈ H

and F ∈ V

;

2) E is coercive; i.e., there exists a C > 0 such that |E(v, v)| ≥ Ckvk2H for

all v ∈ V.

Then by Lions’s representation theorem there exists u ∈ H such that

E(u, v) = F (v) (4.3) for all v ∈ V. For arbitrary w ∈ V and ψ ∈ D(0, T ) we take v(t) =Rt

0

Rs

0 ψ(r) dr ds w.

It follows from (4.3) that ¨

u(t) + B(t) ˙u(t) + A(t)u(t) = f(t)

in L2(0, T ; V

). This identity applied to (4.3) implies that

η( ˙u(0) | ˙v(0))V + η(u(0) | v(0))V = η(u1| ˙v(0))V + η(u0| v(0))V

for all v ∈ V. Hence u(0) = u0and ˙u(0) = u1. This means that u ∈ MR(V, V, H)

is a solution of (4.2).

It remain to prove properties 1) and 2). Again, 1) is obvious and we focus on 2). Let v ∈ V. For ǫ ∈ (0, α) set

R:= ηk ˙v(0)k2V + ηkv(0)k2V +12eλTh (α − ǫ)k ˙v(T )k2V + (λα − ˙ M2 ǫ )kv(T )k 2 V i −12b1(0, ˙v(0), ˙v(0)) − λ 2a1(0, v(0), v(0)) − Re a1(0, v(0), ˙v(0)).

By the V -boundedness of a1and b1 we have

R 12eλTh(α − ǫ)k ˙v(T )k2V + (λα − M˙ǫ2)kv(T )k2Vi+ (η −M

2 )k ˙v(0)k

2

(16)

+ (η −λM2 )kv(0)k2V − Mk ˙v(0)kVkv(0)kV.

Young’s inequality yields

R≥ C1 h k ˙v(T )k2V + kv(T )k 2 V + k ˙v(0)k 2 V + kv(0)k 2 V i

for some C1>0 provided λ and η are sufficiently large. Now

Re E(v, v) = Z T 0 eλtk¨vk2H dt + Z T 0 eλtRe b1(t, ˙v, ¨v) dt + Z T 0 eλtRe b2(t, ˙v, ¨v) dt + Z T 0 eλtRe a1(t, v, ¨v) dt + Z T 0 eλtRe a2(t, v, ¨v) dt + η( ˙v(0) | ˙v(0))V + η(v(0) | v(0))V.

We apply assertion (iii) of Lemma 4.2, it follows that

Re E(v, v) ≥ Z T 0 eλtk¨vk2H dt + Z T 0 eλtRe b2(t, ˙v, ¨v) dt + Z T 0 eλtRe a2(t, v, ¨v) dt +1 2(αλ − 2 ˙M− 2M) Z T 0 eλt k ˙vk2V dt +1 2(λ(αλ − ˙M) − ˙M) Z T 0 eλtkvk2V dt + R.

Thus V -boundedness of a2 and b2and Young’s inequality yield

Re E(v, v) ≥ Ckvk2H,

for some C > 0 provided that λ and η are sufficiently large. This proves 2). Finally, we have seen that the unique solution u satisfies ˙u(T ) ∈ V but we may replace in the previous arguments [0, T ] by [0, t] for any fixed t ∈ (0, T ) and obtain ˙u(t) ∈ V .

2- Piecewise Lipschitz-continuous forms. Suppose now that the forms a1 and

b1satisfy assumption d). We may replace in the first step the interval [0, T ] by [τi−1, τi]. There exists a solution ui ∈ H1(τi−1, τi; V ) ∩ H2(τi−1, τi; H) of the

equation ¨

v(t) + B(t) ˙u(t) + A(t)u(t) = f(t) a.e. t ∈ [τi−1, τi],

with prescribed ui

i−1), ˙ui(τi−1) in V . We also know from the previous step

that ui

i), ˙ui(τi) ∈ V . Now we can solve the previous equation on [τi, τi+1] and

obtain a solution ui+1 such that ui+1

i) = ui(τi) and ˙ui+1(τi) = ˙ui(τi). We

define u on [0, T ] by u = uion [τ

i−1, τi]. It is easy to check that u ∈ MR(V, V, H)

(17)

5

The Wave Equation

Let H, V be Hilbert spaces such that V ֒→ H. Suppose a: [0, T ] × V × V →d

C is a Lipschitz-continuous, symmetric, V-bounded and quasi-coercive non-autonomous form. We denote again by A(t) the operator associated with a(t) on V

and by A(t) the part of A(t) in H. We introduce the maximal regularity space

MR(V, H, V) := L2(0, T ; V ) ∩ H1(0, T ; H) ∩ H2(0, T ; V′) for the second order Cauchy problem. We have the following result.

Theorem 5.1. There exists a unique solution u ∈ MR(V, H, V) of the

non-autonomous second order Cauchy problem

( ¨

u(t) + A(t)u(t) = f(t) t-a.e. u(0) = u0,˙u(0) = u1

(5.1)

for every u0 ∈ V , u1 ∈ H and f ∈ L2(0, T ; H). Moreover u(t) ∈ V for all t∈ [0, T ].

Note that by [DL88, p. 579], for every u ∈ MR(V, H, V), ˙u can be viewed

as a continuous function from [0, T ] into the interpolation space (H, V)

1 2. In

particular, ˙u(0) is well defined and ˙u(0) ∈ V.

We start with the following lemma. Here ˙a(t, ., .) denotes the derivative of

t7→ a(t, ., .).

Lemma 5.2. Let v ∈ H2(0, T ; V ) with ˙v(T ) = 0. Then

(i) Z T 0 eλtRe(¨v| ˙v)H dt = λ 2 Z T 0 eλtk ˙vk2H dt −1 2k ˙v(0)k 2 H (ii) Z T 0 eλtRe a(t, v, ˙v) dt = λ 2 Z T 0 eλta(t, v, v) dt −1 2 Z T 0 eλt˙a(t, v, v) dt +1 2a(T, v(T ), v(T )) − 1 2a(0, v(0), v(0))

Proof. For the first part we calculate the formula eλtk ˙vk2H˙ = −λeλtk ˙vk2H+ 2eλtRe(¨v,˙v)H.

For (ii) we use Lemma 2.1 and the product rule to obtain

eλta(t, v, v)˙ = −λeλta(t, v, v) + 2eλtRe a(t, v, ˙v) + eλt˙a(t, v, v). Now the Lemma follows by integrating over t.

Proof of Theorem 5.1. First we prove existence of a solution. We define the

Hilbert space H := {u ∈ L2(0, T ; V ) ∩ H1(0, T ; H) : u(0), u(T ) ∈ V } with norm

kukHsuch that kuk 2 H:= kuk 2 L2(0,T ;V )+k ˙uk 2 L2(0,T ;H)+ku(0)k 2 V+ku(T )k 2 V and the

pre-Hilbert space V := {v ∈ H2(0, T ; V ) : ˙v(T ) = 0} with norm k.k

V := k.kH. Further we define E : H × V → C by E(u, v) := − Z T 0 ˙u (eλt˙v)˙ H dt

(18)

+ Z T

0

eλta(t, u, ˙v) dt + a(0, u(0), v(0))

and for u0∈ V , u1∈ H and f ∈ L2(0, T ; V) we define F : V → C by

F(v) := Z T

0

eλthf, ˙vi dt + a(0, u0, v(0)) + (u1| ˙v(0))H.

As in the previous sections, we use Lions’s representation Theorem. Suppose that the assumptions of Lions’s Theorem are satisfied. Then there exists a

u∈ H such that

E(u, v) = F (v) (5.2) for all v ∈ V. For the particular choice of v(t) := ψ(t)w where ψ ∈ D(0, T ) and

w∈ V we obtain from (5.2) that

Z T 0 h ˙u, wi(eλtψ˙(t))˙ dt =Z T 0 hf − Au, wieλtψ˙(t) dt.

This implies that ˙u ∈ H1(0, T ; V

), hence u ∈ MR(V, H, V

) and that ¨

u(t) + A(t)u(t) = f(t) t-a.e. (5.3) Following the proof of Lemma 1 and Theorem 2 in [DL88, p. 571 and 575] we can integrate by parts in the first term of E(u, v) to obtain

E(u, v) = h ˙u(0), ˙v(0)i + Z T 0 eλth¨u, ˙vi dt + Z T 0 eλta(t, u, ˙v) dt + a(0, u(0), v(0)) = Z T 0 eλthf, ˙vi dt + a(0, u0, v(0)) + (u1| ˙v(0))H

where we used the identity (5.2). This together with (5.3) implies that h ˙u(0), ˙v(0)i + a(0, u(0), v(0)) = a(0, u0, v(0)) + (u1| ˙v(0))H.

Since v ∈ V was arbitrary this shows that u(0) = u0and ˙u(0) = u1.

Next we check the assumptions of Theorem 2.4. Assumption 1) is again easy to verify. Let v ∈ V, then integration by parts yields to

|E(v, v)| ≥ Re E(v, v) = k ˙v(0)k2H+ Z T 0 eλtRe(¨v| ˙v)H dt + Z T 0 eλtRe a(t, v, ˙v) dt + a(0, v(0), v(0)).

Thus Lemma 5.2 applied to the first and second integral and Young’s inequality shows that Re E(v, v) ≥ 12k ˙v(0)k2H+ λ 2 Z T 0 eλtk ˙vk2H dt + λ 2 Z T 0 eλta(t, v, v) dt

(19)

−12 Z T 0 eλt˙a(t, v, v) dt + 1 2a(T, v(T ), v(T )) + 1 2a(0, v(0), v(0)) ≥ Ckvk2H

for some C > 0 if λ is large enough. Note that we can choose C depending only on the coercivity, V -boundedness, Lipschitz constant of the form and on T .

Uniqueness: Let u ∈ MR(V, H, V) be a solution of (5.1) where f = 0 and u0 = u1 = 0. We have to show that u = 0. Fix r ∈ [0, T ] and define vr(t) :=

RT t 1[0,r]u(s) ds. Then vr∈ H 1(0, T ; V ) with v r(r) = 0 and ˙vr= −1[0,r]u. We obtain 0 = 2 Z T 0 Reh¨u, v ri dt + 2 Z T 0 Re a(t, u, vr) dt = 2 Z r 0 Z r t Reh¨u(t), u(s)i ds dt − 2 Z r 0 Re a(t, ˙vr, vr) dt = 2 Z r 0 Z s 0 Reh¨u(t), u(s)i dt ds − 2 Z r 0 Re a(t, ˙vr, vr) dt = 2 Z r 0 Reh Z s 0 ¨ u(t) dt, u(s)i ds − Z r 0 (a(t, vr, vr))˙− ˙a(t, vr, vr) dt = 2 Z r

0 Reh ˙u, ui ds + a(0, v

r(0), vr(0)) − Z r 0 ˙a(t, vr, vr) dt ≥ ku(r)k2H+ αkvr(0)k2V − ˙M Z r 0 kv rk2V dt. We set w(r) := vr(0) =R r 0u(s) ds ∈ L 2(0, T ; V ). Then w(r) − w(t) = v r(t) and αkw(r)k2V ≤ ˙M Z r 0 kw(r) − w(t)k 2 V dt ≤ 2r ˙Mkw(r)k 2 V + 2 ˙M Z r 0 kw(t)k 2 V dt.

Let 0 < r0 < 2 ˙αM and set Cr0 := α − 2r0M >˙ 0, then for every r ∈ [0, r0] we

have kw(r)k2V ≤ 2 ˙M C −1 r0 Z r 0 kw(t)k 2 V dt.

We conclude by Gronwall’s lemma that w(r) = 0 for all r ∈ [0, r0], hence u = 0

on [0, r0]. Now we may proceed inductively to obtain u = 0 on [0, T ].

Remark 5.3. If we add a V × H-bounded perturbation to a as in Section 4, we

can still prove existence in Theorem 5.1. But for the uniqueness we have to assume additionally that this perturbation is also H × V -bounded.

Remark 5.4. Let B(t) be bounded operators on H with kB(t)kL(H)≤ MB for

a.e. t ∈ [0, T ]. We consider the wave equation (

¨

u(t) + B(t) ˙u(t) + A(t)u(t) = f(t) t-a.e.

u(0) = u0, ˙u(0) = u1

(5.4)

Then for u0 ∈ V , u1 ∈ H and f ∈ L2(0, T, V) there exists a solution u ∈ MR(V, H, V

) to (5.4). The proof is the same as above, one has only to change

E(u, v) into E(u, v) := − Z T 0 ˙u (eλt˙v)˙ H dt

(20)

+ Z T 0 eλt(B(t) ˙u | ˙v)H dt + Z T 0 eλta(t, u, ˙v) dt + a(0, u(0), v(0)).

The uniqueness of u is however not clear except if the map t 7→ B(t) is Lipschitz. If this later condition is satisfied one can use similar ideas as in [DL88, p. 686] to prove uniqueness. The proof for uniqueness in Theorem 5.1 is similar to that of [DL88, p. 673].

6

Applications

In this section we give applications of our results. We consider two problems, one is linear and the second one is quasi-linear.

I) Laplacian with time dependent Robin boundary conditions.

Let Ω be a bounded domain of Rd with Lipschitz boundary Γ. Denote by σ be

the (d − 1)-dimensional Hausdorff measure on Γ. Let

β1, β2: [0, T ] × Γ → R

be bounded measurable functions which are Lipschitz continuous w.r.t. the first variable, i.e.,

|βi(t, x) − βi(s, x)| ≤ M|t − s| (i = 1, 2) (6.1)

for some constant M and all t, s ∈ [0, T ], x ∈ Γ. We consider the symmetric forms a, b: [0, T ] × H1(Ω) × H1(Ω) → R defined by a(t, u, v) = Z Ω∇u∇v dx + Z Γ β1(t, .)uv dσ. (6.2) and b(t, u, v) = Z Ω∇u∇v dx + Z Γ β2(t, .)uv dσ. (6.3)

respectively. The forms a, b are H1(Ω)-bounded and quasi-coercive. The first

statement follows readily from the continuity of the trace operator and the boundedness of β. The second one is a consequence of the inequality

Z Γ|u| 2 dσ ≤ ǫkuk2H1 (Ω)+ cǫkuk 2 L2 (Ω), (6.4)

which is valid for all ǫ > 0 (cǫ is a constant depending on ǫ). Note that (6.4)

is a consequence of compactness of the trace as an operator from H1(Ω) into L2(Γ, dσ), see [Nec67, Chap. 2 § 6, Theorem 6.2].

Let A(t) be the operator associated with a(t, ., .) and B(t) the operator asso-ciated with b(t, ., .). Note that the part A(t) in H := L2(Ω) of A(t) is interpreted

as (minus) the Laplacian with time dependent Robin boundary conditions

(21)

Here we use the following weak definition of the normal derivative. Let v ∈

H1(Ω) such that ∆v ∈ L2(Ω). Let h ∈ L2(Γ, dσ). Then ∂νv = h by definition

ifR Ω∇v∇w + R Ω∆vw = R Γhwdσ for all w ∈ H

1(Ω). Based on this definition,

the domain of A(t) is the set

D(A(t)) = {v ∈ H1(Ω) : ∆v ∈ L2(Ω), ∂νv+ β1(t)v|Γ= 0},

and for v ∈ D(A(t)) the operator is given by A(t)v = −∆v.

Maximal regularity on H for the first order Cauchy problem associated with

A(t) was proved in [ADLO13]. Here we study the second order problem. By Theorem 4.1, the damped wave equation

     ¨ u(t) − ∆ ˙u(t) − ∆u(t) = f(t) u(0) = u0, ˙u(0) = u1∈ H1(Ω)

∂ν( ˙u(t) + u(t)) + β2(t, .) ˙u(t) + β1(t, .)u(t) = 0 on Γ

has a unique solution u ∈ MR(V, V, H) = H2(0, T ; L2(Ω)) ∩ H1(0, T ; H1(Ω))

whenever f ∈ L2(0, T, L2(Ω)).

Indeed, Theorem 4.1 implies that there exists u ∈ MR(V, V, V

) with u(0) =

u0, ˙u(0) = u1and

u, v)H+ b(t, ˙u, v) + a(t, u, v) = (f, v)H (6.5)

for all v ∈ V and all t ∈ [0, T ] \ N, where N is a Lebesgue null set. Let

t∈ [0, T ] \ N, then for the special choice v ∈ D(Ω) we obtain that (6.5) implies ¨

u(t) − ∆ ˙u(t) − ∆u(t) = f(t). This together with (6.5) and the above definition

of the normal derivative shows

∂ν( ˙u(t) + u(t)) + β2(t, .) ˙u(t) + β1(t, .)u(t) = 0 on Γ. II) A quasi-linear problem.

Let Ω be a bounded open set of Rdand let H be the real-valued Hilbert space

L2(Ω, dx) and V be a closed subspace of H1(Ω) which contains H1

0(Ω). If V 6= H01(Ω) we assume that Ω is a Lipschitz domain to ensure that the embedding

of V in H is compact. This latter property is always true for V = H1 0(Ω) for

any bounded domain Ω.

For g, h ∈ L2(0, T ; H) we define the forms a

g,h, bg,h: [0, T ] × V × V → R by ag,h(t, u, v) = d X k,j=1 Z Ω ajk(t, x, g, h)∂ku∂jv dx and bg,h(t, u, v) = d X k,j=1 Z Ω bjk(t, x, g, h)∂ku∂jv dx.

We assume that the coefficients ajk, bjk: [0, T ] × Ω × R × R → R are uniformly

bounded on [0, T ] × Ω × R × R by a constant M > 0 and satisfy the usual ellipticity condition d X k,j=1 ajk(t, x, y, z)ξkξj≥ η|ξ|2, d X k,j=1 bjk(t, x, y, z)ξkξj ≥ η|ξ|2

(22)

for a.e. (t, x) ∈ [0, T ] × Ω and all y, z ∈ R, ξ ∈ Rd. Here η > 0 is a constant. Moreover we assume that ajk(t, x, ., .), bjk(t, x, ., .) are continuous for a.e. (t, x).

We denote by Ag,h(t) and Bg,h(t) the associated operators.

Given u0 ∈ V , u1 ∈ H and f ∈ L2(0, T ; V′) the second order Cauchy

problem

( ¨

u(t) + Bg,h(t) ˙u(t) + Ag,h(t)u(t) = f (t) t-a.e.

u(0) = u0, ˙u(0) = u1

(6.6) has a unique solution ug,h∈ MR(V, V, V′) by Theorem 3.1. Moreover, by

Propo-sition 3.3 there exists C > 0 and 0 < T0≤ T depending only on M and η such

that the solution of (6.6) on [0, T0] satisfies the estimate

kug,hkMR T0(V,V,V′)≤ C h ku0kV + ku1kH+ kfkL2(0,T 0;V′) i . (6.7) Note that C and T0 are independent of g and h. We want to show that the

quasi-linear problem (

¨

u(t) + Bu, ˙u(t) ˙u(t) + Au, ˙u(t)u(t) = f (t) t-a.e.

u(0) = u0,˙u(0) = u1

(6.8)

has a solution u in MR(V, V, V). We define the mapping S : H1(0, T ; H) → H1(0, T ; H) by Sg := ug, ˙g. Note that by (6.7) and the fact that C is independent

of g and h, Im(S) is a bounded subset of MR(V, V, V). Moreover by

Aubin-Lions lemma, MR(V, V, V) is compactly embedded into H1(0, T ; H). Therefore,

if S is continuous then we can apply Schauder’s fixed point theorem to obtain

u ∈ H1(0, T ; H) such that Su = u. Thus u is also in MR(V, V, V

) and u ∈ Im(S). Hence u is a solution of (6.8).

It remains to prove that S is continuous. Let gn → g in H1(0, T ; H) and

set un := Sgn. Since a sequence converges to a fixed element u if and only

if each subsequence has a subsequence converging to u we may deliberately take subsequences. Since L2(0, T ; H) is isomorphic to L2((0, T ) × Ω) we may

assume (after taking a sub-sequence) that gn → g and ˙gn → ˙g for a.e. (t, x).

Furthermore since the sequence un is bounded in MR(V, V, V′) we may assume

(after taking a sub-sequence) that un → u in H1(0, T ; H) and un ⇀ u in

MR(V, V, V). Hence a

jk(t, x, gn,˙gn) → ajk(t, x, g, ˙g) and bjk(t, x, gn,˙gn) →

bjk(t, x, g, ˙g) for a.e. (t, x). Now the equality un= Sgn means that

h¨un,viL2(0,T ;V),L2(0,T ;V )+ d X j,k=1 (∂j˙un| bjk(t, x, gn,˙gn)∂kv)L2(0,T ;H) + d X j,k=1 (∂jun| ajk(t, x, gn,˙gn)∂kv)L2 (0,T ;H)= hf, viL2 (0,T ;V),L2 (0,T ;V )

for all v ∈ L2(0, T ; V ) and u

n(0) = u0, ˙un(0) = u1. By the dominated

con-vergence theorem ajk(t, x, gn,˙gn)∂kv→ ajk(t, x, g, ˙g)∂kv in L2(0, T ; H).

More-over un ⇀ u in MR(V, V, V) implies that ∂jun ⇀ ∂ju and ∂j˙un ⇀ ∂j˙u in

L2(0, T ; H). Thus taking the limit for n → ∞ yields

h¨u,viL2(0,T ;V),L2(0,T ;V )+

d

X

j,k=1

(23)

+

d

X

j,k=1

(∂ju| ajk(t, x, g, ˙g)∂kv)L2(0,T ;H)= hf, viL2(0,T ;V),L2(0,T ;V )

for all v ∈ L2(0, T ; V ) and u(0) = u

0, ˙u(0) = u1. Note that for the initial

condition we have used that MR(V, V, V

) ֒→ C1([0, T ]; H) ∩ C([0, T ]; V ), see

(3.3). This is equivalent to Sg = u. Hence S is continuous.

References

[ADLO13] W. Arendt, D. Dier, H. Laasri and E.M. Ouhabaz. Maximal regular-ity for evolution equations governed by non-autonomous forms. Preprint 2013, available at http://arxiv.org/abs/1303.1166.

[BCS08] C.J.K. Batty, R. Chill and S. Srivastava. Maximal regularity for second order non-autonomous Cauchy problems. Studia Math. 189 (2008) no 3, 205-223.

[DL88] R. Dautray and J.L. Lions. Analyse Mathématique et Calcul Numérique

pour les Sciences et les Techniques. Vol. 8, Masson, Paris, 1988.

[DL92] R. Dautray and J.L. Lions. Mathematical Analysis and Numerical

Meth-ods for Science and Technology. Vol. 5. Evolution problems I,

Springer-Verlag, Berlin, 1992.

[Lio59] J.L. Lions. Problemi misti nel senso di Hadamard classici e generalizzati.

Milan Journal of Mathematics 28 (1958), 147-188.

[Lio61] J.L. Lions. Equations Différentielles Opérationnelles et Problèmes aux

Limites. Springer-Verlag, Berlin, Göttingen, Heidelberg, 1961.

[Nec67] J. Necaš. Les Méthodes Directes en Théorie des Equations Elliptiques. Masson Paris 1967.

[OS10] E. M. Ouhabaz and C. Spina. Maximal regularity for nonautonomous Schrödinger type equations. J. Differential Equations 248 (2010),1668-1683.

[Paz83] A. Pazy. Semigroups of Linear Operators and Applications to Partial

Differential Equations. Springer-Verlag, Berlin, 1983.

[Sho97] R. E. Showalter. Monotone Operators in Banach Space and Nonlinear

Partial Differential Equations. Mathematical Surveys and Monographs.

American Mathematical Society, Providence, RI, 1997.

Dominik Dier, Institute of Applied Analysis, University of Ulm, 89069 Ulm,

Germany,

dominik.dier@uni-ulm.de

El Maati Ouhabaz, Institut de Mathématiques (IMB), Univ. Bordeaux, 351,

cours de la Libération, 33405 Talence cedex, France, Elmaati.Ouhabaz@math.u-bordeaux1.fr

Références

Documents relatifs

Ouhabaz, Maximal regularity for non-autonomous evolution equa- tions governed by forms having less regularity, Arch. Spina, Maximal regularity for non-autonomous Schr¨odinger

Ouhabaz, Maximal regularity for non-autonomous evolution equa- tions governed by forms having less regularity, Arch. Spina, Maximal regularity for non-autonomous Schrödinger

We prove maximal L p -regularity results and other regularity properties for the solutions of the above equation under minimal regularity assumptions on the forms and the

We have compared the similarity among (i) the representative mapping sets from the harmonisation (see Section 2), (ii) the UMLS-based reference mappings of the track, and (iii)

ABSTRACT - We give a uniqueness result concerning the minimizers of the func- tional proposed by Mumford and Shah in order to study the problem of image segmentation

It is expected that the result is not true with = 0 as soon as the degree of α is ≥ 3, which means that it is expected no real algebraic number of degree at least 3 is

The method to construct a diffusion on a finitely ramified fractal is now well-known (cf. [19], [16], [12]): one constructs a Dirichlet form on the fractal X as the limit of

We also found that as the number of ratings of a movie increases its rating similarity approaches to its Wikipedia content similarity whereas its rating similarity diverges away